• Aucun résultat trouvé

Asymptotic Hecke algebras and involutions

N/A
N/A
Protected

Academic year: 2021

Partager "Asymptotic Hecke algebras and involutions"

Copied!
12
0
0

Texte intégral

(1)

http://dx.doi.org/10.1090/conm/610/12156

Asymptotic Hecke algebras and involutions

G. Lusztig

Introduction and statement of results

0.1. In [11], a Hecke algebra module structure on a vector space spanned by

the involutions in a Weyl group was defined and studied. In this paper this study is continued by relating it to the asymptotic Hecke algebra introduced in [6]. In particular we define a module over the asymptotic Hecke algebra which is spanned by the involutions in the Weyl group. We present a conjecture relating this module to equivariant vector bundles with respect to a group action on a finite set. This gives an explanation (not a proof) of a result of Kottwitz [3] in the case of classical Weyl groups, see 2.5. We also present a conjecture which realizes the module in [11] terms of an ideal in the Hecke algebra generated by a single element, see 3.4.

0.2. Let W be a Coxeter group with set of simple reflections S and with length

function l : W → N.

Let A = Z[v, v−1] where v be an indeterminate. We set u = v2. Let A be

the subring Z[u, u−1] of A. Let H (resp. H) be the free A-module (resp. free

A-module) with basis ( ˙Tw)w∈W (resp. (Tw)w∈W). We regard H (resp. H) as an

associative A-algebra (resp. A-algebra) with multiplication defined by ˙TwT˙w =

˙

Tww if l(ww) = l(w) + l(w), ( ˙Ts+ 1)( ˙Ts− u) = 0 if s ∈ S (resp. TwTw = Tww if

l(ww) = l(w) + l(w), (Ts+ 1)(Ts− u2) = 0 if s∈ S). For y, w ∈ W let Py,wbe the

polynomial defined in [2]. For w ∈ W let ˙cw = v−l(w)



y∈W ;y≤wPy,w(u) ˙Ty ∈ H,

cw = u−l(w)



y∈W ;y≤wPy,w(u2)Ty ∈ H, see [2]. Let y ≤LR w, y ∼LR w, y ∼L w

be the relations defined in [2]. We shall write , ∼ instead of ≤LR,∼LR. The

equivalence classes in W under ∼ (resp. ∼L) are called two-sided cells (resp. left

cells).

For x, y, z ∈ W we define ˙hx,y,z ∈ A, hx,y,z ∈ A by ˙cx˙cy =



z∈W ˙hx,y,z˙cz,

cxcy=



z∈Whx,y,zcz. Note that hx,y,z is obtained from ˙hx,y,z by the substitution

v→ u.

0.3. In this subsection we assume that W is a Weyl group or an (irreducible)

affine Weyl group. From the definitions we have:

(a) if ˙hx,y,z = 0 (or if hx,y,z= 0) then z  x and z  y.

For z∈ W there is a unique a(z) ∈ N such that ˙hx,y,z∈ va(z)Z[v−1] for all x, y∈ W 2010 Mathematics Subject Classification. Primary 20G99.

Supported in part by National Science Foundation grant DMS-0758262.

c

2014 American Mathematical Society

(2)

and ˙hx,y,z∈ v/ a(z)−1Z[v−1] for some x, y∈ W . (See [5].) Hence for z ∈ W we have

hx,y,z∈ ua(z)Z[u−1] for all x, y∈ W and hx,y,z∈ u/ a(z)−1Z[u−1] for some x, y∈ W .

For x, y, z ∈ W we have ˙hx,y,z = γx,y,z−1va(z) mod va(z)−1Z[v−1], γx,y,z−1 ∈ Z;

hence we have hx,y,z= γx,y,z−1ua(z) mod ua(z)−1Z[u−1].

(b) If x, y∈ W satisfy x  y then a(x) ≥ a(y). Hence if x ∼ y then a(x) = a(y). (See [5].)

LetD be the set of distinguished involutions of W (a finite set); see [6, 2.2]). Let J be the free abelian group with basis (tw)w∈W. For x, y ∈ W we set

txty =



z∈W γx,y,z−1tz ∈ J (the sum is finite). This defines an associative ring

structure on J with unit element 1 =d∈Dtd (see [6, 2.3]).

0.4. Let∗ : W → W (or w → w∗) be an automorphism of W such that S∗= S,

2= 1. Let I

={w ∈ W ; w∗ = w−1}; if ∗ = 1 this is the set of involutions in W .

Let M be the freeA-module with basis (aw)w∈I. Following [11] for any s∈ S we

define anA-linear map Ts: M → M by

Tsaw= uaw+ (u + 1)asw if sw = ws∗> w;

Tsaw= (u2− u − 1)aw+ (u2− u)asw if sw = ws∗< w;

Tsaw= asws∗ if sw= ws∗> w;

Tsaw= (u2− 1)aw+ u2asws∗ if sw= ws∗< w.

The following result was proved in the setup of 0.3 in [11] and then in the general case in [10].

(a) These linear maps define an H-module structure on M .

Let H = A ⊗AH, M = A ⊗AM . We regard H as a subring of H and M as a

subgroup of M by ξ → 1 ⊗ ξ. Note that the H-module structure on M extends naturally to an H-module structure on M .

Let (Aw)w∈I be the A-basis of M defined in [11, 0.3]. (More precisely, in

[11, 0.3] only the case where W is a Weyl group and∗ = 1 is considered in detail; the other cases are briefly mentioned in [11, 7.1]. A definition, valid in all cases is given in [10, 0.3].)

0.5. In the remainder of this section we assume that W is as in 0.3. For x ∈ W , w, w ∈ I we define fx,w,w ∈ A by cxAw =



w∈I∗fx,w,wAw. The

following result is proved in 1.1:

(a) For x∈ W , w, w∈ Iwe have fx,w,w = βx,w,wv2a(w )

mod v2a(w)−1Z[v−1] where βx,w,w ∈ Z. Moreover, if βx,w,w = 0 then x ∼ w ∼ w.

Let M be the free abelian group with basis (τw)w∈I∗. For x∈ W , w ∈ I we set txτw=



w∈Iβx,w,wτw. (The last sum is finite: if βx,w,w = 0 then fx,w,w = 0

and we use the fact that cxAw is a well defined element of M .) We have the

following result.

0.6 Theorem. The bilinear pairing J× M → M defined by tx, τw→ txτw is

a (unital) J -module structure on M.

The proof is given in§1.

0.7 Notation. Let C be the field of complex numbers. For any abelian group A we set A = C⊗ A.

1. Proof of Theorem 0.6

1.1. In this section we assume that W is as in 0.3. For any x, w∈ W we have

˙cx˙cw˙cx∗−1=



(3)

(a) Hx,w,w =



y∈W ˙h(x, w, y) ˙h(y, x∗−1, w).

From the geometric description of the elements Aw in [11] one can deduce that:

(b) if x ∈ W and w, w ∈ I then there exist elements Hx,w,w+ , Hx,w,w−  of

N[v, v−1] such that Hx,w,w = Hx,w,w+ + Hx,w,w−  and fx,w,w = Hx,w,w+  − Hx,w,w− .

(This fact has been already used in [11, 5.1] in the case where W is finite and

∗ = 1.) Let n ∈ Z, x ∈ W and w, w ∈ I; from (b) we deduce:

(c) If the coefficient of vn in Hx,w,w is 0 then the coefficient of vn in fx,w,w

is 0.

(d) If the coefficient of vn in H

x,w,w is 1 then the coefficient of vn in fx,w,w

is±1.

We can now prove 0.5(a). Setting a0= a(w) we have

Hx,w,w =  y∈W ;wy ˙h(x, w, y) ˙h(y, x∗−1, w) =  y∈W ;a(y)≤a0) ˙h(x, w, y) ˙h(y, x∗−1, w) =  y∈W ;a(y)≤a0

(γx,w,y−1va(y)+ lin.comb.of va(y)−1, va(y)−2, . . . )

× (γy,x∗−1,w−1va0+ lin.comb.of va0−1, va0−2, . . . )

= 

y∈W ;a(y)=a0

γx,w,y−1γy,x∗−1,w−1)v2a0+ lin.comb.of v2a0−1, v2a0−2, . . . .

Using this and (c) we deduce that

fx,w,w = βx,w,wv2a0+ lin.comb.of v2a0−1, v2a0−2, . . . )

where βx,w,w ∈ Z and that if βx,w,w = 0 then γx,w,y−1 = 0, γy,x∗−1,w−1 = 0 for

some y∈ W . For such y we have x ∼ w ∼ y−1, y∼ x∗∼ w−1, see [6, 1.9]. We see that 0.5(a) holds.

The proof above shows also:

(e) if βx,w,w = 0 then for some y ∈ W we have γx,w,y−1 = 0, γy,x∗−1,w−1 = 0.

We show:

(f) If x∈ W and w, w ∈ I satisfy fx,w,w = 0 then w w and w x.

Using (c) we see that Hx,w,w = 0 hence for some y ∈ W we have ˙h(x, w, y) = 0

and ˙h(y, x−1, w)= 0. It follows that y  x, y  w, w y and (f) follows.

1.2. Let x, y ∈ W, w ∈ I. We show that (txty)τw = tx(tyτw) or equivalently

that, for any w∈ I,

(a)y∈Wγx,y,y−1βy,w,w =z∈I

∗βx,z,wβy,w,z

From the equality (cxcy)Aw= cx(cyAw) in M we deduce that

(b)y∈W hx,y,yfy,w,w =



z∈Ifx,z,wfy,w,z.

Let a0 = a(w). In (b), the sum over y can be restricted to those y such that fy,w,w = 0 hence (by 1.1(f)) such that w  y (hence a(y)≤ a0); the sum over z

can be restricted to those z such that fx,z,w = 0 hence (by 1.1(f)) such that w z

(hence a(z)≤ a0). Thus we have



y∈W ;a(y)≤a0

hx,y,yfy,w,w =



z∈I;a(z)≤a0

(4)

Using 0.5(a) this can be written as follows 

y∈W ;a(y)≤a0

(γx,y,y−1v2a(y )

+ lin.comb.of va(y)−1, va(y)−2, . . . )

× (βy,w,wv2a0+ lin.comb.of v2a0−1, v2a0−2, . . . )

= 

z∈I;a(z)≤a0

(βx,z,wv2a0+ lin.comb.of v2a0−1, v2a0−2, . . . )

× (βy,w,zv2a(z)+ lin.comb.of v2a(z)−1, v2a(z)−2, . . . )

that is,  y∈W ;a(y)=a0 γx,y,y−1v2a0βy,w,wv2a0+ lin.comb.of v4a0−1, v4a0−2, =  z∈I;a(z)=a0 βx,z,wv2a0βy,w,zv2a0)+ lin.comb.of v4a0−1, v4a0−2, . . . .

Taking the coefficient of v4a0 in both sides we obtain

 y∈W ;a(y)=a0 γx,y,y−1βy,w,w =  z∈I;a(z)=a0 βx,z,wβy,w,z.

Now, if γx,y,y−1 = 0 then a(y) = a0 and if βx,z,w = 0 then a(z) = a0. Hence we

deduce  y∈W γx,y,y−1βy,w,w =  z∈I βx,z,wβy,w,z.

This proves (a).

1.3. Let w∈ I. We show that 1τw= τwor equivalently that, for any w ∈ I,

(a)d∈Dβd,w,w = δw,w

Let d0 be the unique element ofD contained in the left cell of w−1 (see [6, 1.10]).

If βd,w,w = 0 with d ∈ D then using 1.1(e) we can find y ∈ W such that γd,w,y−1 =

0, γy,d∗,w−1 = 0. (Note that d∗ ∈ D.) Using [6, 1.8,1.4,1.9,1.10] we deduce

γw,y−1,d= 0, γw−1,y,d∗ = 0 and y = w, y = w, d = d0, γw,y−1,d= γw−1,y,d∗ = 1.

Thusd∈Dβd,w,w = βd0,w,w and



y∈W

γd0,w,y−1γy,d∗,w−1 = γd0,w,w−1γw,d∗0,w−1δw,w = δw,w.

Thus the coefficient of v2a(w)in H

d0,w,w is δw,w. Using 1.1(c),(d) we deduce that

the coefficient of v2a(w)in f

d0,w,w is±δw,w that is, βd0,w,w =±δw,w. Thus

(b) 1τw= (w)τw

where (w) =±1. Applying 1 =d∈Dtdto both sides of (b) and using the identity

(11)τw = 1(1τw) that is 1τw = 1(1τw) we obtain (w)τw = 1((w)τw) = (w)2τw

hence (w)2= (w). Since (w) =±1 it follows that (w) = 1. This completes the

proof of (a). Theorem 0.6 is proved.

1.4. For any two-sided cell c of W let Jc(resp. Mc) be the subgroup of J (resp.

M) generated by {tx; x ∈ c} (resp. {τw; w ∈ c ∩ I∗}. Note that Jc is a subring

of J with unit element 1c =



d∈D∩cτd and J =⊕cJc (direct sum of rings). We

have M = ⊕cMc. From the last sentence in 0.5(a) we see that JcMc ⊂ Mc and

JcMc = 0 and for any two sided cells c= c. It follows that the J -module structure

(5)

1.5. For any left cell λ of W such that λ = λ∗let Jλ∩λ−1(resp. ∩λ−1) be the

subgroup of J (resp. M) generated by {tx; x∈ λ∩λ−1} (resp. {τw; w∈ λ∩λ−1∩I∗}.

Note that Jλ∩λ−1is a subring of J with unit element tdwhere d is the unique element

ofD ∩ λ. Since λ = λ∗we have d = d∗. If x∈ λ ∩ λ−1, w∈ λ ∩ λ−1∩ I, w ∈ Iare such that βx,w,w = 0 then w ∈ λ ∩ λ−1∩ I. (Indeed, as we have seen earlier, we

have γx,w,y−1 = 0, γy,x∗−1,w−1 = 0 for some y ∈ W . For such y we have x ∼Lw−1,

w ∼L y, y−1 ∼L x−1, y ∼L x∗, x∗−1 ∼L w, w−1 ∼L y−1, see [6, 1.9]. Hence

y∈ λ, y−1∈ λ w−1∈ λ, w∈ λ∗= λ, so that w∈ λ ∩ λ−1, as required.) It follows that the J -module structure onM restricts for any λ as above to a Jλ∩λ−1-module

structure on∩λ−1. Now if d ∈ D−λ, w ∈ λ∩λ−1∩I, w ∈ Ithen βd,w,w = 0

so that tdτw= 0. (Indeed, assume that βd,w,w = 0. Then, as we have seen earlier

we have γd,w,y−1 = 0 for some y ∈ W . We then have d∼Lw−1, see [6, 1.9], hence

d ∈ λ, contradiction.) Since 1τw = τw it follows that tdτw = tw. We see that the

Jλ∩λ−1-module structure onMλ∩λ−1 is unital.

2. Γ-equivariant vector bundles

2.1. Let V ec be the category of finite dimensional vector spaces over C.

Let Γ be a finite group and let X be a finite set with a given action (a Γ-set). A Γ-equivariant C-vector bundle (or Γ-v.b.) V on X is just a collection of objects Vx∈ V ec (x ∈ X) with a given representation of Γ on ⊕x∈XVx such that

gVx= Vgx for all g∈ Γ, x ∈ X. We say that Vxis the fibre of V at x. Now X× X

is a Γ-set for the diagonal Γ-action. Let C0 be the category whose objects are the

Γ-v.b. on X× X. For V ∈ C0 let Vx,y ∈ V ec be the fibre of V at (x, y); for g ∈ Γ

letTg: Vx,y → Vgx,gy be the isomorphism given by the equivariant structure of V .

For V, V ∈ C0we define the convolution VV ∈ C0by

(VV)x,y =⊕z∈XVx,z⊗ Vz,y

for all x, y in X with the obvious Γ-equivariant structure. For V, V, V ∈ C0 we

have an obvious identification (VV)V= V(VV). Let Cδ ∈ C0 be the

Γ-v.b. given by (Cδ)x,x= C for all x∈ X and (Cδ)x,y = 0 for all x= y in X (with

the obvious Γ-equivariant structure). For V ∈ C0 we have obvious identifications CδV = V = V Cδ. Define σ : X× X → X × X by σ(x, y) = (y, x). For V ∈ C0

we set Vσ = σV that is Vσ

x,y = Vy,x for all x, y in X. For V, V ∈ C0 we have an

obvious identification (VV)σ= VVσ. Note that is compatible with direct

sums in both the V and V factor. Hence if K(C0) is the Grothendieck group ofC0

then  induces an associative ring structure on K(C0) with unit element defined

by Cδ; moreover, V → Vσ induces an antiautomorphism of the ring K(C0). Thus K(C0) is an associative C-algebra with 1.

2.2. Let C be the category whose objects are pairs (U, κ) where U ∈ C0 and κ : U → U∼ σ is an isomorphism inC0 (that is a collection of isomorphisms κx,y :

Ux,y→ Uy,x for each x, y∈ X such that κgx,gyTg=Tgκx,y for all g∈ Γ, ξ, y ∈ X);

it is assumed that κy,xκx,y = 1 : Ux,y → Ux,y for all x, y∈ X.

For V ∈ C0 we define an isomorphism ζ : V ⊕ Vσ → (V ⊕ Vσ)σ= Vσ⊕ V by a⊕ b → b ⊕ a. We have (V ⊕ Vσ, ζ)∈ C and V → (V ⊕ Vσ, ζ) can be viewed as

functor Θ : C0 → C. Let K(C) be the Grothendieck group of C and let K(C) be

the subgroup of K(C) generated by the elements of the form Θ(V ) with V ∈ C0.

(6)

in [9, 11.1.5] which applies to a category with a periodic functor.) Note that if (U, κ)∈ C then (U, −κ) ∈ C and (U, κ) + (U, −κ) = 0 in ¯K(C).

For V ∈ C0, (U, κ)∈ C we define V ◦ (U, κ) ∈ C by V ◦ (U, κ) = (V UVσ, κ)

where for x, y in X,

κx,y :⊕z,z∈XVx,z⊗ Uz,z⊗ Vy,z → ⊕z,z∈XVy,z ⊗ Uz,z⊗ Vx,z

maps a⊗ b ⊗ c (in the z, z summand) to c⊗ κ(b) ⊗ a (in the z, z summand). Now

let V, V∈ C0 and (U, κ)∈ C. We have canonically

(V ⊕ V)◦ (U, κ) = V ◦ (U, κ) ⊕ V◦ (U, κ) ⊕ Θ(V UVσ).

Moreover, we have canonically V ◦ Θ(V) = Θ(VVVσ). For V, V ∈ C

0 and

(U, κ) ∈ C we have an obvious identification (VV ) ◦ (U, κ) = V◦ (V ◦ (U, κ)). For (U, κ)∈ C we have an obvious identification Cδ◦ (U, κ) = (U, κ). We see that

◦ defines a (unital) K(C0)-module structure on ¯K(C) (but not on K(C)). Hence

¯

K(C) is naturally a (unital) K(C0)-module.

2.3. Note that K(C0) has a Z-basis consisting of the the isomorphism classes

of indecomposable Γ-v.b. V on X× X (these are indexed by a Γ-orbit in X × X and an irreducible representation of the isotropy group of a point in that orbit). Moreover, ¯K(C) has a signed Z-basis consisting of the classes of (V, κ) where V is an indecomposable Γ-v.b. on X× X satisfying V ∼= Vσ and κ is defined up to a sign (so the class of (V, κ) is defined up to a sign).

2.4. Let CΓ be the category of Γ-v.b. on Γ viewed as a Γ-set under

conjuga-tion. An object Y of CΓ is a collection of objects Yg ∈ V ec (g ∈ Γ) with a given

representation of Γ on ⊕g∈ΓYg such that gYg = Yggg−1 for all g, g ∈ Γ, x ∈ X.

For Y, Y ∈ CΓ we define the convolution YY∈ CΓ by

(YY)g=⊕g1,g2∈Γ;g1g2=gYg1⊗ Y

 g2

for all g ∈ Γ with the obvious Γ-equivariant structure. This defines a structure of associative ring with 1 on the Grothendieck group K(CΓ). The unit element is

given by the Γ-v.b. whose fibre at g = 1 is C and whose fibre at any other element is 0. Hence K(CΓ) is an associative C-algebra with 1; by [8, 2.2], it is commutative

and semisimple.

With X,C0,C as in 2.1, for any Y ∈ CG, we define as in [8, 2.2(h)] an object

Ψ(Y ) ∈ C0 by Ψ(Y )x,y = ⊕g∈Γ;x=gyYg (with the obvious equivariant structure).

Now Y → Ψ(Y ) defines a ring homomorphism K(CΓ) → K(C0) and a C-algebra

homomorphism K(CΓ)→ K(C0). By [8, 2.2],

(a) K(C0) is a semisimple C-algebra and the image of the homomorphism K(CΓ)→ K(C0) is exactly the centre of K(C0).

We see that the K(C0)-module structure on ¯K(C) restricts to a K(CΓ)-module

struc-ture on ¯K(C) in which the product of the class of Y ∈ CΓwith the class of (V, κ)∈ C

is the class of (V, κ)∈ C where

Vx,y =⊕g,g∈Γ,z,z∈X;x=gz,y=gzYg⊗ Vz,z⊗ Yg

that is,

(b) Vx,y =⊕g,g∈ΓYg⊗ Vg−1x,g−1y⊗ Yg

and, for x, y in X,

(7)

maps a⊗ b ⊗ c (in the g, g summand) to c⊗ κ(b) ⊗ a (in the g, g summand). It

follows also that ¯K(C) is naturally a K(CΓ)-module.

Now assume in addition that

(c) Γ is an elementary abelian 2-group

and that Y ∈ CΓ is such that for some g0 ∈ Γ, Y |Γ−{g0} is zero and dim Yg0 = 1.

Then (b) becomes

Vx,y = Yg0⊗ Vg0x,g0y⊗ Yg0

Now Yg0⊗Yg0is isomorphic to C as a representation of Γ and Vg0x,g0yis canonically

isomorphic to Vx,y. We see that (V, κ) = (V, κ). Thus Y acts as identity in

the K(CΓ)-module structure of ¯K(C). It follows that if Y is any object of CΓ

then Y acts in the K(CΓ)-module structure of ¯K(C) as multiplication by ν(Y ) =



g∈Γdim Yg. Note that ν defines a ring homomorphism K(CΓ) → Z and a

C-algebra homomorphism K(CΓ)→ C (taking 1 to 1). We see that:

(d) If Γ is as in (c) then for any ξ ∈ K(CΓ), ξ ∈ ¯K(C) we have ξξ = ν(ξ)ξ. In particular, the K(CΓ)-module ¯K(C) is ν-isotypic.

Using this and (a) we see that the first assertion in (e) below holds.

(e) If Γ is as in (c) then the K(C0)-module ¯K(C) is isotypic. Moreover dimCK(C)¯ is equal to |Γ| times the number of Γ-orbits in X.

We now prove the second assertion in (e). By 2.3, dimCK(C) is equal to n¯ Γ,X, the

number of indecomposable Γ-v.b. on Xτ X (up to isomorphism) such that V ∼= Vσ. For such V there exists a unique Γ-orbitO on X such that Vx,y= 0 implies x ∈ O

and y ∈ O. Hence nΓ,X =



OnΓ,O where O runs over the Γ-orbits in X. This

reduces the proof to the case where X is a single Γ-orbit. Let H be the isotropy group in Γ of some point in X; this is independent of the choice of point since Γ is commutative. Now any (x, y) ∈ X × X is in the same Γ-orbit as (y, x). (In-deed, we can find g ∈ Γ such that y = gx. Then (x, y) is in the same orbit as (gx, gy) = (y, g2x) = (y, x) since g2= 1.) For a given Γ-orbitO in Xτ X the

num-ber of indecomposable Γ-v.b. on X× X (up to isomorphism) with support equal to O is the number of characters of characters of the isotropy group of any point in the orbit which is H. Thus nΓ,X is equal to|H| times the number of Γ-orbits in X× X that is to |H| × |Γ/H| = |Γ|. This proves (e).

2.5. In this subsection we assume that W is an irreducible Weyl group and ∗ = 1. Let c be a two-sided cell of W such that for z ∈ c we have a(z) = 11 (if W

is of type E7) and a(z) = 11, a(z) = 26 (if W is of type E8). Let Γ be the finite

group associated to c in [8, 3.15]. For each left cell λ in c let Γλ be the subgroup of

Γ associated to λ in [8]. Let X =λ(Γ/Γλ) (λ runs over the left cells in c). Note

that Γ acts naturally on X. For any λ as above let Cλ be the Γ-v.b. on X× X

which is C at any point of form (x, x), x∈ Γ/Gλ, and is zero at all other points.

LetC0 be defined in terms of this X. The following statement was conjectured in

[8, 3.15] and proved in [1]:

(a) There exists a isomorphism φ : Jc→ K(C∼ 0) which carries the basis {tx; x∈

c} onto the canonical basis 2.3 of K(C0) and is such that for any left cell λ in c, φ(td) (where D ∩ λ = {d}) is the class of Cλ.

The isomorphism φ has the following property conjectured in [7, 10.5(b)] in a closely related situation.

(8)

(b) Let x ∈ c and let φ(tx) be the class of the indecomposable Γ-v.b. V on

Xτ X. Then φ(tx−1) is the class of ( ˇV )σ where ˇV is the dual Γ-v.b. to V .

(As R. Bezrukavnikov pointed out to me, (b) follows immediately from (a).) Next we note the following property.

(c) ˇV ∼= V for any Γ-v.b. on X× X.

It is enough to show that for any (x, y)∈ X ×X, the stabilizer of (x, y) in Γ (that is the intersection of a Γ-conjugate of Γλwith a Γ-conjugate of Γλ where λ, λare two

left cells in c) is isomorphic to a Weyl group (hence its irreducible representations are selfdual). This can be verified from the explicit description of the subgroups Γλ

in [8].

Using (b),(c) we see that φ has the following property.

(d) Let x ∈ c and let φ(tx) be the class of the indecomposable Γ-v.b. V on

Xτ X. Then φ(tx−1) is the class of Vσ.

I want to formulate a refinement of (a).

(e) Conjecture. There exists an isomorphism of abelian groups ψ :Mc→ ¯K(C)

(C is defined in terms of X) with the following properties:

-if w∈ c ∩ I and φ(tw) = V (an indecomposable Γ-v.b. such that V ∼= Vσ, see

(d)) then ψ(τw) = (V, κ) for a unique choice of κ : V → V∼ σ;

-the Jc-module structure on Mc corresponds under φ and ψ to the K(C0 )-module structure on ¯K(C).

Now let Jc = C⊗ Jc, Mc = C⊗ Mc. Note that Jc is a semisimple algebra, see

[8, 1.2, 3.1(j)]. Assuming that (e) holds we deduce:

(f) If Γ is as in 2.4(c) then the Jc-moduleMc is isotypic. Moreover, dimCMc is equal to |Γ| times the number of left cells contained in c.

(Note that the number of Γ-orbits on X is equal to the number of left cells contained in c.)

Now if W is of classical type, then Γ is as in 2.4(c) and (f) gives an explanation for the known structure of the W -module obtained from M for u = 1 (a consequence of the results of Kottwitz [3]); this can be viewed as evidence for the conjecture (e). (In this case, the second assertion of (f) was already known in [4, 12.17].)

Here we use the following property which can be easily verified for any Weyl group.

(g) Let Mc (resp. Mc−c) be the A submodule of M spanned by {Ax; x 

y for some y ∈ c} (resp. {Ax; x  y for some y ∈ c, x /∈ c}). The decomposition

pattern of the (semisimple) Jc-moduleMcis the same as the decomposition pattern of the (semisimple) C(v)⊗AH-module C(v)A(Mc/Mc−c); in particular if the first module is isotypic then so is the second module.

One can show, using results in [6, 2.8, 2.9], that this property also holds when W is replaced by an affine Weyl group and c by a finite two-sided cell in that affine Weyl group.

3. A conjectural realization of the H-module M

3.1. Let H = Q(u)AH (an algebra over Q(u)) and let M = Q(u)AM .

We regard H as a subset of H and M as a subset of M• by ξ → 1 ⊗ ξ. The H-module structure on M extends in an obvious way to an H•-module structure on M•. Let ˆH be the vector space consisting of all formal (possibly infinite) sums 

x∈WcxTx where cx ∈ Q(u). We can view H as a subspace of ˆH in an obvious

(9)

way to a H-module structure on ˆH. We set

X= 

x∈W ;x∗=x

u−l(x)Tx∈ ˆH.

Let M = H•Xbe the H-submodule of ˆH generated by X. In this section we will give a conjectural realization of the H•-module M•in terms of M.

We write S = {si; i ∈ I} where I is an indexing set. For any sequence

i1, i2, . . . , ik in I we write i1i2. . . ik instead of si1si2. . . sik ∈ W .

3.2. In this subsection we assume that W is of type A2,∗ = 1 and S = {s1, s2}.

We set

X= (−u)−3T121+ u−2T12+ u−2T21+ u−1T1+ u−1T2+ 1, X1= (1 + u)−1(T1− u)X = (u− 1)(u−3T121+ u−2T12+ u−1T1), X2= (1 + u)−1(T2− u)X = (u− 1)(u−3T121+ u−2T21+ u−1T2),

X121= T1X2= T2X1= (u− 1)((u−1+ u−2− u−3)T121+ u−1T12+ u−1T21).

Clearly, X, X1, X2, X121form a basis of M. In the H•-module M• we have a1= (u + 1)−1(T1− u)a, a2= (u + 1)−1(T2− u)a, a121= T1a2= T2a1.

We see that we have a (unique) isomorphism of H-modules M→ M∼ such that

X→ a, X1→ a1, X2→ a2, X121→ a121.

3.3. In this subsection we assume that W is of type A1,∗ = 1 and S = {s1}.

We set X = u−1T1+ 1, X1 = (u− 1)u−1T1. Clearly, X, X1 form a basis of M.

In the H•-module M• we have a1 = (u + 1)−1(T1− u)a. We see that we have a

(unique) isomorphism of H-modules M→ M∼ such that X→ a, Xs→ a1. 3.4. We return to the setup in 3.1. Based on the examples in 3.2, 3.3 we state:

(a) Conjecture. There exists a unique isomorphism of H•-modules η : M→ M∼ such that X→ a.

By 3.3, 3.2, conjecture (a) is true when W is of type A1, A2(wth∗ = 1). It can be

shown that it is also true when W is a dihedral group (any ∗) or of type A3 (any ∗).

Assuming that (a) holds we set Xw= η−1(aw) for any w∈ I.

3.5. We describe below the elements Xw for various w∈ I∗when W is of type

A3and∗ = 1. We write S = {s1, s2, s3} (s1s3= s3s1). Xhas been described in 3.4;

X1= (u− 1)(u−1T1+ u−2T12+ u−2T13+ u−3T121+ u−3T123+ u−3T132 + u−4T1213+ u−4T1232+ u−4T1321+ u−5T13213+ u−5T12132+ u−6T121321); X3= (u− 1)(u−1T3+ u−2T32+ u−2T13+ u−3T323+ u−3T321+ u−3T132 + u−4T3231+ u−4T3212+ u−4T1323+ u−5T13213+ u−5T32312+ u−6T121321); X2= (u− 1)(u−1T2+ u−2T21+ u−2T23+ u−3T121 + u−3T323+ u−3T213+ u−4T1213+ u−4T3231+ u−4T2132 + u−5T32312+ u−5T12132+ u−6T121321); X13= (u− 1)2(u−2T13+ u−3T132+ u−4T1321+ u−4T1323+ u−5T13213+ u−6T121321);

(10)

X121= (u− 1)(u−1T12+ u−1T21+ (u−1+ u−2− u−3)T121+ u−2T123 + u−2T213+ (u−2+ u−3− u−4)T1213+ u−3T1323+ u−3T2132 + (u−3+ u−4− u−5)T12132+ u−4T13213+ u−4T21321 + (u−4+ u−5− u−6)T121321); X323= (u− 1)(u−1T32+ u−1T23+ (u−1+ u−2− u−3)T323+ u−2T321 + u−2T213+ (u−2+ u−3− u−4)T3213+ u−3T1321+ u−3T2132 + (u−3+ u−4− u−5)T32132+ u−4T13213+ u−4T21323 + (u−4+ u−5− u−6)T121321); X2132= (u− 1)2(u−2T213+ u−3T2132+ u−4T21321 + u−4T21323+ u−4T12321+ (u−4+ u−5− u−6)T121321); X13213= (u− 1)(u−1T132+ u−1T123+ u−1T321+ (u−1+ u−2− u−3)T1321 + u−2T3213+ (u−1+ u−2− u−3)T1323+ u−2T1213+ u−2T2132 + (u−2+ u−3− u−4)T21323+ (u−2+ u−3− u−4)T21321 + (2u−2+ u−3− 2u−4)T13213+ (u−2+ 2u−3− u−4− 2u−5+ u−6)T121321); X213213= (u− 1)2(u−2T1213+ u−2T2132+ u−2T2321 + (u−2+ u−3− u−4)T21323+ (u−2+ u−3− u−4)T21321 + (u−2− u−4)T13231+ (u−2+ u−3− u−4− u−5+ u−6)T121321).

3.6. We describe below the elements Xw for various w ∈ I when W is an

infinite dihedral group and ∗ = 1. We write S = {s1, s2}. X has been described

in 3.4; X1= (u− 1)(u−1T1+ u−2T12+ u−3T121+ u−4T1212+ . . . ); X2= (u− 1)(u−1T2+ u−2T21+ u−3T212+ u−4T2121+ . . . ); X121= (u− 1)(u−1T12+ u−2T121+ u−3T1212+ u−4T12121+ . . . ); X212= (u− 1)(u−1T21+ u−2T212+ u−3T2121+ u−4T21212+ . . . ); X12121= (u− 1)(u−1T121+ u−2T1212+ u−3T12121+ u−4T121212+ . . . ); X21212= (u− 1)(u−1T212+ u−2T2121+ u−3T21212+ u−4T212121+ . . . ); X1212121= (u− 1)(u−1T1212+ u−2T12121+ u−3T121212+ u−4T1212121+ . . . ); X2121212= (u− 1)(u−1T2121+ u−2T21212+ u−3T212121+ u−4T2121212+ . . . ); . . .

(11)

3.7. Assume that Conjecture 3.4(a) holds for W,∗. From the examples above

we see that it is likely that the elements Xw (w ∈ I) are formal Z[u−1]-linear

combinations of elements Tx (x∈ W ). In particular the specializations (Xw)u−1=0

are well defined Z-linear combinations of Tx (x ∈ W ). From the example above

it appears that there is a well defined (surjective) function π : W → I such that (Xw)u−1=0=



x∈π−1(w)Tx. We describe the sets π−1(w) in a few cases with∗ = 1.

If W is of type A1 we have π−1(∅) = {∅}, π−1(1) ={1}. If W is of type A2 we have π−1(∅) = {∅}, π−1(1) ={1}, π−1(2) ={2}, π−1(121) ={12, 21, 121}. If W is of type A3 we have π−1(∅) = {∅}, π−1(1) ={1}, π−1(2) ={2}, π−1(3) ={3}, π−1(13) ={13}, π−1(121) ={12, 21, 121}, π−1(323) ={32, 23, 323}, π−1(2132) ={213}, π−1(13213) ={132, 123, 321, 1321, 1323}, π−1(121321) ={1213, 2132, 2321, 21323, 21321, 13231, 121321}. If W is infinite dihedral we have

π−1(∅) = {∅}, π−1(1) ={1}, π−1(2) ={2}, π−1(121) ={12}, π−1(212) ={21},

π−1(12121) ={121}, π−1(21212) ={212}, π−1(1212121) ={1212},

π−1(2121212) ={2121}, . . .

In each of these examples π is given by the following inductive rule. We have

π(∅) = ∅. If x ∈ W is of the form x = six with i∈ I, x ∈ W , l(x) > l(x) so that

π(x) can be assumed known, then

π(x) = siπ(x) if siπ(x) = π(x)si> π(x),

π(x) = siπ(x)si if siπ(x)= π(x)si> π(x),

π(x) = π(x) if π(x)si< π(x).

In each of the examples above the following holds: if x∈ W , w = π(x) ∈ I, then

l(w) = l(x) + l(x−1w). We expect that these properties hold in general.

References

[1] Roman Bezrukavnikov, Michael Finkelberg, and Victor Ostrik, On tensor categories attached

to cells in affine Weyl groups. III, Israel J. Math. 170 (2009), 207–234, DOI

10.1007/s11856-009-0026-9. MR2506324 (2011a:20006)

[2] David Kazhdan and George Lusztig, Representations of Coxeter groups and Hecke algebras, Invent. Math. 53 (1979), no. 2, 165–184, DOI 10.1007/BF01390031. MR560412 (81j:20066) [3] Robert E. Kottwitz, Involutions in Weyl groups, Represent. Theory 4 (2000), 1–15

(elec-tronic), DOI 10.1090/S1088-4165-00-00050-9. MR1740177 (2000m:22014)

[4] George Lusztig, Characters of reductive groups over a finite field, Annals of Mathematics Studies, vol. 107, Princeton University Press, Princeton, NJ, 1984. MR742472 (86j:20038) [5] George Lusztig, Cells in affine Weyl groups, Algebraic groups and related topics

(Kyoto/Nagoya, 1983), Adv. Stud. Pure Math., vol. 6, North-Holland, Amsterdam, 1985, pp. 255–287. MR803338 (87h:20074)

[6] George Lusztig, Cells in affine Weyl groups. II, J. Algebra 109 (1987), no. 2, 536–548, DOI 10.1016/0021-8693(87)90154-2. MR902967 (88m:20103a)

[7] George Lusztig, Cells in affine Weyl groups. IV, J. Fac. Sci. Univ. Tokyo Sect. IA Math. 36 (1989), no. 2, 297–328. MR1015001 (90k:20068)

[8] G. Lusztig, Leading coefficients of character values of Hecke algebras, The Arcata Conference on Representations of Finite Groups (Arcata, Calif., 1986), Proc. Sympos. Pure Math., vol. 47, Amer. Math. Soc., Providence, RI, 1987, pp. 235–262. MR933415 (89b:20087)

(12)

[10] G. Lusztig, A bar operator for involutions in a Coxeter group, Bull. Inst. Math. Acad. Sin. (N.S.) 7 (2012), no. 3, 355–404. MR3051318

[11] George Lusztig and David A. Vogan Jr., Hecke algebras and involutions in Weyl groups, Bull. Inst. Math. Acad. Sin. (N.S.) 7 (2012), no. 3, 323–354. MR3051317

Références

Documents relatifs

We report on a proof-checked version of Stone’s result on the representability of Boolean algebras via the clopen sets of a totally disconnected compact Hausdor↵ space.. Our

Note that we do not give the proof of Theorem 2 since it follows from the proof of Theorem 1 exactly as in the case of the semilinear heat equation 6 treated in [9].. We give

The common point to Kostant’s and Steinberg’s formulae is the function count- ing the number of decompositions of a root lattice element as a linear combination with

The proof of the Milin conjecture here given differs from the argument verified by the Leningrad Seminar in Geometric Function Theory [6] in that it avoids approxima- tion

Seeking to contribute to these applications, this research project proves the Yau Geometric Con- jecture in six dimensions, a sharp upper bound for the number of positive lattice

In the remainder of this paper, we will implicitly assume that all link diagrams are connected, and that a diagram is alternating within each twist region... Thurston [29,

We derive an asymptotic formula which counts the number of abelian ex- tensions of prime degrees over rational function

– We prove a conjecture of Kottwitz and Rapoport which implies a converse to Mazur’s Inequality for all connected split and quasi-split unramified reductive groups.. Our results